The most general density-cubic equation of state ... - ACS Publications

Aug 1, 1982 - K. Hemanth Kumar, Kenneth E. Starling. Ind. Eng. Chem. Fundamen. , 1982, 21 (3), pp 255–262. DOI: 10.1021/i100007a011. Publication Dat...
0 downloads 0 Views 873KB Size
Ind. Eng. Chem. Funahm. 1982, 21, 255-262

255

The Most General Density-Cubic Equation of State: Application to Pure Nonpolar Fluids K. Hemanth Kumar and Kenneth E. Starling' School of Chemlcal Engineerlng and Meterlals Science, University of Oklahoma, Norman, Oklahoma 73019

Thermodynamic properties of pure nonpolar fluids are predicted accurately, over wide ranges of temperature and pressure conditions, by the most general density-cubic equation of state. Of particular merit is the accurate predlctkn of liquid densities and bw-temperature vapor pressures. The development of the temperature dependence and the generalization of the equation of state using hemodynamic property data for normal paraffin hydrocarbons, methane through ndecane, are presented herein. The generalized equation of state predicts vapor pressure and vapor and liquid densities to within 1% (average)of the experimental values for these fluids. The equation of state prediction of thermodynamic propertles of 32 fluids is compared with the predictions by a popular density-cubic equation namely, the Peng-Robinson equation and a popular BWR-type equation of state. According to these comparisons the most general density-cubic equation is s u p e r b to the Peng-Robinson equation and is on par with the BWR-type equation of state.

Introduction Many attempts have been made over the years to describe the thermodynamic behavior of real fluids via equations of state. These equations of state have achieved varying degrees of success, enabling us to divide them into three separate classes. In the first class, we have the equations of state which are cubic in density. A few of the more popular density-cubic equations are the van der Waals equation (1873), the Redlich-Kwong equation (19491, the Soave equation (1972),and the Peng-Robinson equation (1976). The density-cubic equations of state give reasonable descriptions of the thermodynamic behavior of real fluids, with each equation generally being more accurate in the chronological order of appearance in the literature. The Beattie-Bridgeman equation (1928), the Benedict-Webb-Rubin equation (1940),and the Modified Benedict-Webb-Rubin equation (Starling, 1973) are popular examples of the second class of equations of state. They are noncubic in density and provide a good description of the thermodynamic behavior of real fluids for both vapor and liquid phases. In the third class of equations are the nonanalytic equations of state which are highly constrained for specific fluids (Goodwin, 1977) and give a highly accurate description of real fluid behavior. In most industrial design situations as well as research measurements of derived properties, the unknown variable is density, whereas the easily measurable properties pressure and temperature are known. Consequently, the first class of equations, namely the density-cubic equations, is of particular interest since they provide an analytical solution for the density, as compared to the more complicated noncubic and nonanalytic equations of state, which require time consuming iterative procedures to solve for the density. The presently available popular density-cubic equations of state such as the Soave and the Peng-Robinson equations provide good descriptions of real fluid behavior in the two-phase region and in the gas phase, but in the compressed liquid region they lack by far the accuracy levels attainable using the second class of equations of state. When we look at the form of the density-cubicequations in the chronological order of appearance in the literature, we find that in general the more recent equations have more complex density dependence (when expressed in a pressure explicit form) than their predecessors. For ex0198-4313/82/1021-0255$01.25/0

ample, the Redlich-Kwong equation has more density dependence than the van der Waals equation, as seen below.

p = - -PR T 1 - pb

up2

(van der Waals)

pRT

p=--(1 - Pb)

(1 + Pb)

(Redlich-Kwong)

Similarly,the Peng-Robinson equation of state has more density dependence than the Redlich-Kwong equation

p = - -PRT 1 - Pb

U('I?P2

(1 + 2bp - b2p2)

(Peng-Robinson)

In general, the overall performance in fluid properties prediction is greatly enhanced when using the PengRobinson equation as compared to the Redlich-Kwong equation and the Redlich-Kwong equation in turn is better than the van der Waals equation. Thus, though the temperature dependence of each equation is different, it can be projected that at a particular temperature a higher density dependence leads to a more accurate equation of state. Continuing in the same vein, it can be stated that the most density dependence (in terms of pressure) that can be introduced into a density-cubic equation will in turn lead to the most accurate cubic equation of state. This fact is very important because if the most general density-cubic equation of state can provide an accuracy level comparable to the second class of equations of state for all fluid states it becomes highly desirable in situations where repetitive calculations for the density are required due to its inherent advantages. The derivation of the most general density-cubic equation of state has been presented recently by the authors (Kumar and Starling, 1980). The equation of state is expressed as 1 + d1p + d2p2 Z= (1) 1 d3p + d4p2+ d5p3

+

Equation 1 can be written in a computationally more convenient form by expressing the denominator as a product of a first-order term and a quadratic function by setting dl = As, d2 = A2,d3 = A3 - A I ,d4 = A4 - A1A3, and d5 = -AIAI 0 1982 American Chemical Society

256

*

Ind. Eng. Chem. Fundam., Vol. 21, No. 3, 1982

0.246

where Z is the compressibility factor and pr is the reduced density (plp,) and Al through A, are density independent parameters which can be temperature and composition dependent. Development of a Provisional Temperature Dependence for the Equation of State The requirement that the equation of state be cubic in density restricts the density dependence, but the temperature dependence is open for analysis, with practical application of the equation of state being the only criterion restricting the order of the temperature dependence. The evaluation of polynomials consumes very little computing time as compared to iterative search procedures and thus adding more terms to a polynomial in temperature does not for practical purposes increase the time required for an analytical solution of the cubic equation. In the development of the temperature dependence of the equation of state, the goal was to attain a level of accuracy comparable to the second class of equations of state such as the Three Parameter Corresponding States Mofidied Benedict-Webb-Rubin equation of state, hereinafter referred to as 3PCS-MBWR (Brul6 et al., 1979). In order to determine the form of the temperature dependence, the expressions for the second virial coefficient and third virial coefficient were obtained from eq 2 by expanding it into a polynomial in density. The second and third reduced virial coefficients are given by the following equations B = A5 - A3 + A1 (3) C = A2 - A4 + A1A3 - (A3 - AJA5 + (A3 - AJ2 (4) where B and C are the second and third virial coefficients, respectively. It is known (Tsonopoulos, 1974) that the second and third virial coefficients can be well represented by reciprocal temperature expansions. Since all the five parameters AI through A, appear in the expressions for the second and third virial coefficients, reciprocal temperature expansions were chosen for the temperature dependence of the equation of state. To develop the analytical relations for the temperature dependence of the parameters, thermodynamic property values were used for propane along 25 isotherms from Goodwin (1977), covering a wide range of fluid states from a reduced temperature of 0.24 to a reduced temperature of 1.62 and up to a reduced pressure of 17. At each isotherm, values of density, vapor pressure, and the partial derivative of pressure with respect to density (aPf ap)Twere used in multiproperty regression analysis (Cox et al., 1971; Wang et al., 1976) to determine the optimum set of values for the parameters AI through A,. In the next step, the temperature dependence of each of the parameters was determined. The parameters Al and A4 were independent of temperature while the parameters A2,As, and A, had the following temperature dependence. A, = A51 + T, A52

A3 = A31

+ A53 - + A54 - + A55 T; T,3 T,4

A33 A34 + A32 -+ -+ T, T,2 T,3

(5)

(7)

The above temperature dependence was developed using thermodynamic property values for propane, but to pro-

3' 0 23

0.3

0 2

0 1

Acentric f a c t o r ,

3 4

-

Figure 1. Plot of parameter AI vs. acentric factor, w .

vide the basis for the generalization of the equation of state, the parameters Al, A4 and the coefficients of the temperature functions for A5, A2, and A3 were determined for methane, n-heptane, and n-octane, respectively. For each fluid, the thermodynamic properties density, vapor pressure, and enthalpy departure were used in multiproperty regression analysis to determine the optimum set of parameters. For the fluids considered, namely methane, propane, n-heptane, and n-octane, the overall average absolute deviation of predicted density, vapor pressure, and enthalpy departure from experimental values was about 0.5%. This error in properties prediction is quite small, and thus the equation of state is amenable to generalization. Generalization of the Equation of State In the three-parameter corresponding states theory proposed by Pitzer (1955) the compressibility factor, 2, can be expressed in a power series in the acentric factor w , with the expansion truncated after the first order term z = 2, Z1w ... (8)

+

+

20= Zo(Tr, P r ) 21 = Zl(Tr, Pr)

(9) (10)

where T, = Tf T, and P, = PIPc. For simple fluids like argon the acentric factor is zero and the compressibility factor is given by 2,. For other fluids the acentric factor is calculated from the following defining relation given by Pitzer w = -log Pr - 1.000 (11) where P, is the reduced vapor pressure at T, = 0.70. (2, + Z p ) represents the compressibility factor for fluids which deviate from the simple fluid behavior. This theory has been applied to a wide class of fluids, with greatest success for nonpolar fluids. The parameter values obtained for methane, propane, n-heptane, and n-octane were plotted against acentric factor, w as shown in Figures 1 to 5. From Figures 1 and 2 it can be inferred that Al is a function only of acentric factor and A4 is practically a constant. The temperature dependent parameters A2, A3, and A, have essentially linear dependence on the acentric factor at each reduced temperature. With knowledge of the general functional dependence of the parameters from Figures 1 to 5, the analytical dependence was determined through regression analysis of thermodynamic property data. Experimental density, vapor pressure, and enthalpy departure values for the normal straight chain hydrocarbons methane, ethane, propane, n-butane, n-pentane, n-hexane, n-heptane, n-

Ind. Eng. Chem. Fundam., Vol. 21, No. 3, 1982 257 -0.182

-0.186

5.0

*

A4

,*.

-0.192

4.0'

*

-0.196

0.1

0

-

*

I

0.2

Acentric factor,

-

2.0

-

0.4

0.3

A2 j ,

Figure 2. Plot of parameter A, vs. acentric factor, w.

/< Tr = 0.3

3.0

8

* * *

L

0.5

-

U

1.0-*

0.

-1.

-1.0 O

L

n

w

%

*

0 . 1-

-

0.2

0.6 0 ,

0.3

n

0.4

Acentric factor, w

Figure 4. Plot of the parameter A2 vs. acentric factor, w , at various

-2.

reduced temperatures.

-3. A5

o.6 0.5 -4

t

Tr = 0 . 3

n n

0

1 1

B

-

B

t

0.4 A3

-5. 0.3

0.2

-6.

0

-

@

Y

* .*

!I.

U

B B

L'

@

(4 0.5 e

nfi

70.

a 0.e

009 1.c

-I

2.0 41; u

*-

I

I

V. I

0

0.1

0.2

Acentric factor,

Figure 3. Plot of the parameter A5 vs. acentric factor, w , at various reduced temperatures. octane, n-nonane, and n-decane were used simultaneously in multiproperty regression analysis to arrive at the following generalized equation of state.

where

+

2+

s> (-az +

Al = al

A4 = -al

T,

0.4

u)

Figure 5. Plot of the parameter A3 vs. acentric factor, w, at various

Acentric factor,

A2 = ( a ,

0.3

+T:

103~:

reduced temperatures.

where y is a parameter referred to herein as the orientation parameter. The orientation parameter was set equal to the acentric factor for the regression to determine the parameters a, through a12. The value of the acentric factor, w, for a particular fluid depends on the source from which it is obtained. For example, the value of the acentric factor for methane has been quoted as 0.0072,0.008, and 0.0115, respectively, in three sources (Passut and Danner, 1973; Reid et al., 1977; and Henry and Danner, 1978). This discrepancy is probably due to the fact that the accuracy of the value of the acentric factor depends upon the accuracy of the vapor pressure value at a reduced temperature of 0.7 and the accuracy of the critical temperature and the critical pressure of the fluid. In general, the vapor pressure at the reduced temperature of 0.7 is not measured, and hence this vapor pressure value is either interpolated from values at other temperatures or it is obtained from a vapor pressure equation. In any case, the accuracy in predicted property values obtained from a generalized empirical equation of state depends on the use of the acentric factor value which is consistent with the value used in the development of the equation of state. Thus, in order to make the value of the

258

Ind. Eng. Chem. Fundam., Vol. 21, No. 3,1982

Table

I.

Characterization Parameters for Use With Eq 12 crit oriencrit density, tation fluid temp, K kg-mol/m3 param, y

methane ethane propane is0b ut ane n-butane isopentane n-pentane n-hexane n-heptane n-octane n-nonane n-decane n-undecane n - do de cane n-tridecane n-tetradecane n-pentadecane n-hexadecane n-heptadecane n-octadecane n-nonadecane n-eicosane ethylene propylene hydrogen sulfide nitrogen carbon dioxide benzene naphthalene tetralin quinoline phenanthrene

190.69 305.39 369.80 407.85 425.19 460.37 469.49 507.23 540.29 568.59 594.54 617.54 638.80 658.30 675.80 694.00 707.00 717.00 733.00 745.00 756.00 767.00 283.05 365.04 373.54 126.15 304.15 562.16 748.35 720.15 782.15 873.15

10.0500 6.7566 4.9600 3.9053 3.9213 3.2469 3.2149 2.7167 2.3467 2.0568 1.8421 1.6611 1.5153 1.4026 1.2821 1.2049 1.1363 1.0598 1.0000 0.9430 0.8937 0.8495 8.0653 5.5248 10.5257 11.0992 10.6378 3.8662 2.4186 2.2692 2.4853 1.8066

0.0115 0.0980 0.1520 0.1831 0.1947 0.2246 0.2467 0.2967 0.3467 0.3949 0.4468 0.4880 0.5350 0.5801 0.6213 0.6565 0.7154 0.7860 0.8071 0.8492 0.8965 0.9427 0.0995 0.1475 0.1083 0.0399 0.2109 0.2133 0.2998 0.3196 0.2580 0.4641

Table 11. Generalized Parameters Used in Eq 12

i

1 2 3 4 5 6

ai

0.262524 0.236048 -1.096720 0.042057 0.196828 -1.589350

i

7 8 9 10 11 12

ai

-0.178144 0.549559 0.579249 - 1.297180 0.278052 0.250865

acentric factor compatible with the equation of state, an orientation parameter, y, which in most cases is equal to

the former was determined from regression analysis of thermodynamic property data. The values of the orientation parameter, y, for the fluids studied in this work are presented in Table I along with the other characterization parameters, critical temperature, and critical density. For other fluids, the orientation parameter can either be replaced by the value of the acentric factor, or more preferably it can be determined from regression analysis of thermodynamic property data. The numerical values for the set of generalized parameters a, through u12occurring in eq 12 are presented in Table 11. To use eq 1 2 to calculate a pure fluid thermodynamic property, only the values of the critical temperature, the critical density, and the orientation parameter are required. Equation 12 predicts density and vapor pressure values for methane, ethane, propane, n-butane, n-pentane, nhexane, n-heptane, n-octane, n-nonane, and n-decane covering wide ranges of temperature and pressure conditions to within 1% of the experimental values on the average (Tables IV and V). The enthalpy departure values for methane, ethane, propane, n-butane, n-pentane, nheptane, and n-octane are predicted with 3.95 kJ/kG average absolute deviation (Table VI). Values of the orientation parameter, y, were determined for 22 additional fluids to ascertain the applicability of the equation of state for fluids not used in the generalization. These fluids are isobutane, isopentane, ethylene, propylene, carbon dioxide, hydrogen sulfide, nitrogen, benzene, naphthalene, tetralin, quinoline, phenanthrene, and the straight chain saturated hydrocarbons n-undecane through n-eicosane (Tables IV and V). Comparisons with Other Equations of State The major literature sources from which the experimental data were obtained for use in the development of the most general density-cubic equation of state and for comparisons with other equations of state, are presented in Table 111. The primary objective behind the development of the most general density-cubic equation of state was to provide a cubic equation which could describe the thermodynamic properties of nonpolar fluids, especially liquid densities and low-temperature vapor pressures on par with noncubic equations of state. The noncubic

Table 111. Data References fluid(s) methane ethane propane isobutane n-butane isopentane n-pentane n-hexane n-heptane n-octane n-nonane n-decane n-undecane to n-eicosane ethylene propylene hydrogen sulfide nitrogen carbon dioxide benzene naphthalene tetralin quinoline phenanthrene

major data sources Douslin (1964),Jones (1963),Matthew (1946),Van Itterbeck (1963),Vennix (1966), Yesavage (1968) API (1953),Canjar (1967),Sage (1950) Goodwin (1977),Yesavage (1968) Beattie (1950),Sage (1950),Waxman (1978),Connolly (1962),Zwolinski (1971) API (1953),Canjar (1967), Sage (1950) Schumann (1942),Silberberg (1959),Vohra (1959),Willingham (1945) API (1953),Canjar (1967),Sage (1950) API (1953),Stewart (1954),Canjar (1967) API (1953),Gilliland (1942),Kay (1938), Stuart (1950) API (1953),Felsing (1942),Mundel (1913),Yound (1900),Lenoir (1968)

API (1953) API (1953), Felsing (1942),Mundel(1913), Young (1900),Lenoir (1968) Vargaftik (1975) Canjar (1967),Michels (1936),Rossini (1953),Tickner (1951) Canjar (1967),Michels (1953),Rossini (1953),Tickner (1951) Kay (1953),Lewis (1968),Reamer (1959),West (1948) Canjar (1967),Friedman (1950),Mage (1963),Streett (1968) Canjar (1967),Din (1961)

API (1953) API (1978a),Wilson (1981) API (1978b),Wilson (1981)

API (1979a),Wilson (1981)

MI (1979b),Wilson (1981)

Ind. Eng. Chem. Fundam., Vol. 21, No. 3, 1982 259

Table IV. Comparisons of Density Predictions no. of data fluids points temp range, K methane 41 114.5-623.2 46 70 354 40 116 41 41 41 50 32 17 17 14 14 9 13 13 12 41 61 41 41 41 60 18 17 17 7

ethane

propane isobutane n-butane isopentane n-pentane n-hexane n - he ptan e n-octane n-decane n-dodecane

n-tridecane n-tetradecane n-pentadecane n -hexadecane n-heptadecane n-octadecane

n-nonadecane ethylene propylene hydrogen sulfide

nitrogen carbon dioxide benzene naphthalene

tetralin quinoline

phenanthrene

133.5-427.6 90.0-600.0 185.4-57 3.3 144.3-494.3 124.9-473.1 144.3-51 0.9 177.6-410.9 205.4-510.9 21 6.5-538.7 310.9-510.9 323.2-483.2 343.2-503.2 393.2-523.2 41 3.2-543.2 463.2-543.2 433.2-553.2 453.2-573.2 463.2-573.2 116.5-399.8 227.6-505.4 277.6-444.3 77.6-388.7 243.1-413.1 513.2-613.2 360.0-1000.0 300.0-1000.0 320.0-1000.0 373.2-573.2

equation of state chosen for comparison is the three-parameter corresponding states Modified Benedict-WebbRubin (3PCS-MBWR) equation of state developed by Starling and co-workers (Brul6 et al., 1979). To compare eq 12, a cubic equation, the Peng-Robinson equation of state was chosen as it is one of the most popular cubic equations used in industry today. The comparisons are presented on a relative basis, percent average absolute deviation defined by A.A.D.% =

1

exptl - calcd 5 exptl n i=l

1,

X

100

1

for density and vapor pressure and on an average absolute deviation basis for the enthalpy departure defined by A.A.D. =

n

C Jexptl- calcdli

i=l

where i indicates the ith data point, n is the total number of data points, and exptl and calcd refer to the experimental and calculated values of the data point, respectively. Density. The average absolute deviations of equation of state predictions of the densities of 30 fluids, the number of data points used, and the temperature and pressure ranges of the data are presented in Table IV. In general, other cubic equations have relatively large deviations for liquid densities compared to noncubic equations. This is shown to be true for the Peng-Robinson equation (P-R) in comparison with the three parameter corresponding states modified BWR equation (3PCS-MBWR) in Table IV. However, the most general density-cubic equation (eq 12) is on par with the 3PCS-MBWR equation. Vapor Pressure. In the development of the PengRobinson equation of state, vapor pressure data from the normal boiling point up to the critical point were used (Peng and Robinson, 1976). In the present work, when the

press. range, bar 8.9-160.3 1.0-551.6 1.1-733.3 0.1-344.7 1.O-482.6 0.01-184.4 1.O-689.5 1.O-205.5 1.0-212.5 1.O-16.5 13.8-413.7 0.001-0.871 0.001-0.934 0.01 -0.934 0.01-0.99 0.07-0.696 0.01-0.613 0.02-0.696 0.02-0.5 1.0-138.0 1.1-202.6 6.9-137.9 1.0-616.1 15.2-304.1 25.8-598.1 1.0-20.0 1.o-100.0 1.0-100.0 0.00-0.7

P-R 5.33 5.35 3.84 3.70 3.65 2.52 3.14 1.59 1.47 3.92 4.71 7.80 7.58 7.76 8.95 10.28 13.41 17.14 20.11 3.19 2.51 3.57 4.24 1.99 3.17 1.48 6.82 1.82 8.45

A.A.D. % 3-PCSMBWR 0.65 1.23 1.01 0.87 0.55 1.77 1.15 0.53 0.65 1.18 1.09 0.54 1.66 1.66 2.03 2.1 2 3.57 4.60 2.73 1.58 1.85 0.27 0.65 2.57 2.89 2.91 4.34 0.26

eq12 1.17 1.69 0.62 0.94 0.54 1.25 0.84 0.33 0.53 1.25 0.43 2.22 2.78 2.78 3.97 3.47 4.31 4.32 4.72 1.87 1.44 2.06 1.40 0.83 1.84 2.65 2.66 4.22 4.65

Peng-Robinson equation was used to calculate vapor pressures over the full range of temperature for comparison purposes, the relative errors in the calculated vapor pressures were quite large at temperatures below the normal boiling point. In the range of the normal boiling point to the critical point the vapor pressures are predicted quite accurately, as reported (Peng and Robinson, 1976). Due to the relatively large deviations at low temperatures, the overall deviation for the full range was rather high, and hence vapor pressure comparisons were not appropriate for the Peng-Robinson equation. Vapor pressure predictions for 32 fluids using eq 12 are compared with predictions using the 3PCS-MBWR equation of state in Table V. The range of fluids for which the comparisons are presented include components of natural gas, petroleum, and coal. In most cases low-temperature vapor pressure data were included for comparison. It can be seen from the comparisons presented in Table V that the most general density-cubic equation of state predicts vapor pressures quite well and as accurately as the 3PCS-MBWR equation of state. Enthalpy Departures. Enthalpy departure predictions for eleven fluids are presented in Table VI. Here, eq 12 is compared with the Peng-Robinson equation and the three-parameter corresponding states Modified Benedict-Webb-Rubin equation. According to these comparisons the most general density-cubic equation of state is more accurate than the Peng-Robinson equation and compares quite well with the 3PCS-MBWR equation of state. Conclusions This work presents, for nonpolar fluids, the development of the temperature dependence of the most general density-cubic equation of state. The generalized equation of state, which has been developed thus far in a three-parameter corresponding states framework for nonpolar fluids, can be extended to polar and associative fluids if

260

Ind. Eng. Chem. Fundam., Vol. 21, No. 3, 1982

Table V. Comparisons of Vapor Pressure Predictions

A.A.D. %

fluids

no. of data points

temp range, K

3-PCS-MBWR

eq 1 2

methane ethane propane isobutane n-butane isopentane n-pentane n-hexane n-heptane n-octane n-nonane n-decane n-undecane n-dodecane n-tridecane n-tetradecane n-pentadecane n-hexadecane n-heptadecane n-octadecane n-nonadecane n-eicosane ethylene propylene hydrogen sulfide nitrogen carbon dioxide benzene naphthalene tetralin quinoline phenanthrene

29 46 21 64 38 64 50 53 44 47 24 24 19 22 19 16 14 10 16 15 16 17 36 28 24 19 33 36 15 16 8 12

111.66-190.64 138.71-305.38 120.0-369.80 186.11-407.85 202.59-425.1 6 217.17-460.39 179.60-469.77 219.86-407.87 192.74-531.59 2 18.71-549.82 223.75-452.59 243.50-477.59 347.04-499.82 323.1 5-659.17 343.17-677.17 393.17-695.0 413.17-553.17 463.17-553.17 433.15-583.17 453.17-593.17 463.1 5-613.17 473.15-767.00 133.15-283.05 146.93-364.93 212.93-373.54 88.71-125.93 216.48-304.26 280.30-553.17 441.32-727.61 450.17-71 0.94 533.17-727.61 481.0-727.61

0.68 1.10 0.46 1.61 0.53 0.86 1.03 0.98 0.75 1.16 1.46 0.77 0.44 0.56 0.97 0.58 0.71 0.29 1.51 0.77 1.24

0.98 1.07 0.94 0.49 0.48 0.34 1.19 0.89 0.64 1.18 1.56 1.40 0.56 0.71 0.65 0.67 0.30 0.19 0.32 0.14 0.76 2.77 2.00 1.10 1.21 0.92 0.81 0.51 3.07 2.44 5.58 4.64

2.08 0.69 0.72 0.90 0.76 0.69 6.36 3.63 5.12 9.37

Table VI. Comparisons of Enthalpy Departure Predictions

fluids

no. of data points

temp range, K

press. range, bar

methane ethane propane isobutane n-butane n-pentane n-heptane n-octane ethylene nitrogen carbon dioxide

38 98 39 24 39 39 17 68 38 79 39

116.48-283.15 166.48-427.59 116.48-394.26 311.11-522.22 310.93-494.26 310.93-510.93 539.98-594.0 297.04-588.72 188.71-399.82 88.71-283.1 5 243.1 5-413.15

31.03-1 37.9 13.79-241.3 34.47-137.9 17.24-206.8 13.79-344.7 13.79-689.5 5.43-162.9 13.79-96.53 6.89-1 37.9 13.79-1 72.4 30.41-506.7

adequate characterization is provided for the effects of polarity and hydrogen bonding. This leads to a multiparameter corresponding states framework where the equation of state, and in turn the parameters in the cubic equation, can be expressed as

where /J*is the reduced dipole moment and CY is a measure of associative effects. The generalized equation of state was applied to 32 pure fluids consistingmainly of nonpolar hydrocarbons. Where comparisons have been made, eq 12 is superior to the Peng-Robinson equation of state and is as accurate as the three-parameter corresponding states Modified Benedict-Webb-Rubin equation of state. Finally, the most general density-cubic equation has been shown to perform as well as a noncubic equation such as the BPCS-MBWR equation in the overall fluid states. The equation of state can be extended to mixtures by using conventional mixing

P- R

A.A.D., kJ/kg 3-PCSMBWR

eq12

5.58 3.74 8.37 4.10 3.16 4.16 2.46 6.58 6.58 2.42 7.21

3.23 2.28 3.37 1.77 1.16 1.23 1.72 4.30 4.58 1.12 6.25

4.67 3.88 3.04 2.58 1.45 2.41 2.74 6.69 3.39 1.15 4.99

rules for the characterization parameters. Acknowledgment This work was supported by the University of Oklahoma and the Department of Energy through the Pittsburgh Energy Technology Center, Contract DE-FG2280PC30249. Appendix Expressions for the derived thermodynamic properties for eq 12 were converted into reduced form in terms of the variables pr = p / p c and TI = T/T,. The classical relationships in reduced form for enthalpy departure, entropy departure, internal energy departure, Helmholtz free energy departure, entropy departure, Gibbs free energy departure, and fugacity coefficient for a pure fluid are as follows

Ind. Eng. Chem. Fundam., Vol. 21, No. 3, 1982 201

S-SO -R

-

- i p r ( Z- 1) d In pr - T,~P’(dZ/dT,), 0 d In pr - In P/Z (A.2)

-U-UO RTC -A - A’ RTC

- -T,2S”(dZ/dTr), 0

d In pr

(A.3)

- T,spr(Z- 1) d In pr + T,In P/Z 64.4) 0

di ( i = 1,5) = parameters in eq 1 f = fugacity F,K, Q,X,Y = symbols defined in Appendix C = gibbs free energy H = enthalpy P = pressure R = universal gas constant S = entropy T = temperature U = internal energy 2 = compressibility factor

Greek Letters a = association factor y = orientation parameter

T r s p ’ ( Z- 1)d In pr + Tr In P/Z + Tr(Z - 1) (A.5) 0

In ( f / P ) = S p ’ ( Z- 1) d In pr - In Z + (2 - 1) 0

(A.6)

Subscripts

x p ’ ( Z- 1)d In p,; ~ ” ( d Z / d T r ) pd, In pr

r = reduced property

0

are evaluated from the equation of state, all of the above properties can be determined. The equation of state expressions for the two integrals are given below. Let us define A12 A6A1 A2

F=

+ + Ai2 + A3A1 + A4

K=

A4F - A2

A1 Q = A? - 4A4

X = 1 + A3pr + A4p: 2 + ~r(A3+ Q112) 2 + pr(A3- Q112) Then x p r ( Z- 1) d In pr =

2

Superscripts 0 = ideal gas state

* = reduced

When the integrals

-’(

= dipole moment = molar density w = acentric factor

fi p

’) In X - F In (1 -Alp,)

+ Y(K + A3(F - 1)/2) (A.7)

c = critical property i = parameter index Literature Cited

API Research Project No. 44 “Selected Values of Physlcal and Thermodynamlc Properties of Hydrocarbons and Related Compounds”; Carnegie Press, Camegk Institute of Technology: Pittsburgh, PA, 1953. API Monograph Series, Amerlcan Petroleum Institute Publication No. 705; Washlngton, DC, 1978. API Monograph Series. Amerlcan Petroleum Institute Publication No. 707; Washington, DC, 1978. API Monograph Series, Amerlcan Petroleum Institute Publication No. 708; Washlngton, DC, 1979. API Monograph Series, Amerlcan Petroleum Institute Publication No. 711; Washlngton, DC, 1979. Beattie. J. A.; Marpie, S:Jr.; Edwards, D. G. J. Chem. Phys. 1950, 18, 127. Beattle, J. A.; Brldgeman, 0. C. Roc. Am Aced. Arts Sci. 1928, 63, 229. Benedlct, M.; Webb, 0. B.; Rubln, L. C. J. Chem. Phys. 1940, 8 , 334. Brul, M. R.; Lee, L. L.; Staring, K. E. Chem. Eng. 1979, 86, 155. Canjar, L. N.; Manning, F. S. “Thermodynamlc Properties and Reduced Correlations for Gases”; Gulf Publlshlng Company: Houston, 1967. Connoily, J. F. J. Phys. Chem. 1982, 66, 1082. Cox, K. W.; Bono, J. L.; Kwok, Y. C.; Starling, K. E. Ind. Eng. Chem. Fundam. 1071, 10. 245. Din. F. “Thermodynamic Functions of Gases”, 1; Butterworths: London, 1961. DousHn. D. R.; Hankon, R. H.; Moore, R. T.; McCuHough, J. P. J . Chem. Eng. Data 1984, 9 , 358. Felslng, W. A.; Watson, 0. M. J. Am. Chem. Soc. 1942, 64, 1822. Friedman, A. S.; White, D. J. Am. Chem. Soc. 1950. 72, 3931. Gllliland, E. R.; Parekh, M. D. Ind. Eng. Chem. 1942. 34, 360. (kodwln, R. D. “Provisional Thermodynamlc Functlons of Propane, from 85 to 700 K at Ressure to 700 bar”; NBSIR 77-860, 1977. Henry, W. P.; Danner, R. P. Ind. fng. Chem. Roc8ss Des. D e v . 1978. 17, 373. Jones. M. L; Mage D. T.; Faukner. R. C.; Katz D. L. Chem Eng . Rog Sym . Ser. 1963, 59, 52. Kay, W. B. Ind. Eng. Chem. 1038, 30, 459. Kay, W. B.; Rambosek, 0. M. Ind. fng. Chem. 1059, 45, 221. Kumar, K. H.; Starling, K. E. Ind. Eng. Chem. Fundam. 1980, 19, 128. Len&, J. M.; Robinson, D. R.; Hipkin. H. 0. Preprint No. 20-68. 33rd Midyear Meeting, API DMsron of Reflnlng Philadelphia. May 15, 1988. Lewis, L. C.; Frederlcks, W. J. J. Chem. Eng. Date 1988, 13, 482. Mage, D. T.; Jones, M. L.; Katz, D. L.; Roebuck, J. R. Chem. Eng. Prog. Sym. Ser. 1968, 59, 61. Matthews, C. S.; Hurd, C. 0. Trans. A I C E 1947, 42, 55. Mlchels. A.; De Wer. J.; Niesen, F. Physic8 1936, 3 , 347. Mlchels, A.; Wassenaar, T.; Louwerse, P; Lunbeck, R. J., Woikers, G. J. Physic8 1959, 19. 287. Mundel, C. F. 2.Phys. Chem. W13, 85, 435. Passut. C. A.; Danner, R. P. Ind. fng. Chem. process Des. Dev. 1973, 12, 365. Peng, D. Y.; Robinson, D. B. Ind. Eng. Chem. Fundem. 1976, 15. 59. Pttzer, K. S. J. Am. Chem. Soc. 1955, 77, 3427. Reamer, H. H.; Sage, B. H.; Lacey. W. N. Ind. fng. Chem. 1950. 42, 140. Rediich, 0.; Kwong, J. N. S. Chem. Rev. 1949, 44, 233. Reid, R. C.; Prausnttz, J. M.; Sherwood, T. K. “Properties of Gases and Liquids”. 3rd ed.; McGraw-HIH: New York, 1977. Rosslni, F. D. API Research Project 44; Carnegie Press: Pktsburgh, PA, 1953. Sage, B. H.; Lacey, W. N. “Thermodynamic Properties of the Lighter Paraffin Hydrocarbons and Nitrogen”; API: New York, 1950. Schumann. S. C.; Aston. J. G.; Sagenkahm. M. J . Am. Chem. Soc. 1942, 64, 1039.

.

Nomenclature A = Helmholtz free energy Ai (i = 1,5) = equation of state parameters a = attraction parameter ai (i = 1,12) = generalized parameters in eq 12 B = second virial coefficient b = van der Waals co-volume C = third virial coefficient

.

Ind. Eng. Chem. Fundam. 1982, 2 1 , 262-268

262

Siiberberg, H.; McKetta, J. J.; Kobe, K. A. J. Chem. f n g . Data 1959,4 , 323. Soave, G. Chem. Eng. Sci. 1972,27, 1197. Starling, K. E. "Fluid Thermodynamic Properties for Light Petroleum Systems"; Gulf Publishing Co.: Houston, Texas, 1973. Stewart, D. E.; Sage,B. H.; Lacey, W. N. Ind. Eng. Chem. 1954,46,2529. Streett, W. B.; Staveiey, L. A. K. Adv. Ctyog. Eng. 1988, 13, 363. Stuart, E. 8 . ; Yu, K. T.; Coull, J. Chem. Eng. Prog. lQ50.46,31 1. Tickner. A. W.; Lossing, F. P. J. phvs. Coibid Chem. 1951,55, 733. Tsonopoulos, C. AIChE J. 1974,20,263. van der Waals, J. D. Doctoral Dissertation, LMen, Holland, 1873. Van Ilterbeek, A.; Verbeke, 0.; Staes, K. Physice lB89,29, 742. Vargaftk, N. B. "Tables on the Thermophyslcal Properties of LiquMs and Gases", 2nd ed.;Hemisphere Pub. Co.: Washington, DC, 1975. Vennix, A. J. Ph.D Thesis, Rice University, 1966. Vohra, S. P.; Kobe, K. A. J. Chem. Eng. Data 1959,4 , 329. Wang. J. S.;Van Daei, W.; Starling, K. E. Can. J. Chem. Eng. 1978,5 4 ,

Waxman, M.; Davis, H.; Sengers, L.; Klein, M. "The Equation of State for Isobutane: An Interim Assessment"; Thermophysics Division, NBS, 1978. West, J. R. Chem. Eng. Prog. 1948,4 4 . 287. Wlliingham, C. B.; Taylor, W. J.; Pignocco, J. M.; Rossini, F. D. J. Res. Natl. Bur. Stand. 1945,35, 219. Wilson, G. M.; Johnston, R. H.;Hwang, S.C.; Tsonopouios, C. Ind. Eng. Chem. Process Des. Dev. 1981,20,94. Yesavage, V. F. Ph.D. Thesis, University of Michigan, 1968. Young, S. J. Chem. SOC. 1900, 77,1145. Zwolinski, B. J.; Wiiholt, R. C. "Vapor Pressure and Heats of Vaporization of Hydrocarbons and Related Compounds"; API 44-TRC 101, 1971.

Received f o r review March 23, 1981 Accepted February 24, 1982

241

Gaseous Diffusion in Porous Solids at Elevated Temperatures Ralph T. Yang' Department of Chemical Engineering, State University of New York at Buffalo, Amherst, New York 14260

Rea-Tllng Llu State University of New York at Stony Brook, Stony Brook, New York I1780

The Wicke-Kallenbach (WK) technique was applied to the measurement of the binary gaseous diffusivity in a porous carbon at temperatures up to 700 OC. The severe leaking problem associated with the high-temperatwe application of the technique was solved by electroplating of soft metal layers on the sample. The diffusivity was also independentlycalculated from the reaction rate at a very low conversion through the use of a model. Comparisons of the two values showed that the WK diffusivitiis were substantially lower (by about 60%) than those calculated from reaction conditions. The commonly used expression in modeling and design studies, De, = c2D,, was found to yield values by over an order of magnitude too high in the temperature range of practical interest, e.g., 500-1200 'C. A remedy to this problem is to replace the molecular diffushrity (13,) in the expression by a transition diffusivii (Dt). D,may be approximated by using a hypothetical single pore size.

Introduction Knowledge of the rate of gaseous diffusion in porous solids at elevated temperatures is essential in understanding the kinetics of heterogeneous reactions. For the gas-carbon reactions, information on the pore diffusion at elevated temperatures is desirable in developments of coal conversion processes and of the high-temperature gas-cooled nuclear reactors. In our previous work, the experimental effective diffusion coefficients in several carbonaceous materials were compared with the predicted values using two structural models familar to workers in the field of heterogeneous catalysis (Yang and Liu, 1979). The predicted values were consistently higher than the measured data by a factor ranging from 1.3 to 15.4. The discrepancy was attributed to the highly tortuous path as well as the dead-end pores. That work was limited to temperatures below 100 "C. Diffusion data for the elevated temperature range are scarce, if not nonexistent, in the literature. Roberts and Satterfield (1965) have estimated the effective diffusivities in a carbonaceous material under reaction conditions by using an assigned temperature coefficient and extrapolated from the room temperature data. The WickeKallenbach type (WK) diffusion apparatus has been used for diffusion measurement at elevated temperatures. However, a severe leakage problem exists for the high-temperature work. It is caused by the stronger 0 196-43 1318211021-0262$01.25/0

temperature dependence (near !P/2) of leak through the space between the sample holder and the sample plug, as compared to the temperature dependence of pore diffusion (near P'2).The first experimental demonstration of the severe leakage problem was done by Growcock and coworkers (1977),in which the leaking rate could be an order of magnitude higher than the pore diffusion rate in a graphite sample at 500-800 OC. Golovina reported a temperature coefficient of 1.34 for C02/N2 in a graphite membrane at temperatures ranging from 20 to 600 O C (Golovina, 1969). Nichols measured diffusivities for several binary systems in a graphite at temperatures up to 700 " C (Nichols, 1961). In both cases, the leaking rate was uncertain, and not identified. Makaiho and Takahashi (1963) developed a sealing between graphite and glass by using silver and silver chloride, and they obtained diffusivity data for temperatures up to 200 "C. It is also worth noting that Olsson and McKewan (1966) measured the effective diffusivities under reaction conditions using a technique not related to the WK cell, but the technique is applicable to only a few gas-solid reaction systems. In this work, we solved the leakage problem in the WK cell by using an electroplated metal layer around the graphite sample and obtained binary diffusivities in the temperature range of 27-700 "C. In a parallel but independent study, the effective diffusivity was calculated from the overall reaction rate under reaction conditions, by the 0 1982 American Chemical Society